Translate

Monday, February 29, 2016

The things you can do with Hydrostatic Equilibrium

We are bounded in a nutshell of Infinite Space: Reading #4: The things you can do with Hydrostatic Equilibrium

In the reading of chapter 3 of Maoz’s Astrophysics in a Nutshell, we notice the several references to the Thompson Cross Section, the area of an electron which is hit by energy from the star which further causes radiation to escape the star in many ways. In fact there, are some stars which are no longer supported by the pressure the matter exerts, rather the radiation they produce counteracts the gravity! These stars, of sizes between \( 10 - 300 M_\odot\) are maintained by this radiation pressure, and as such have a different equation describing the relationship of pressure and gravity, but in a way a bit more useful to us. These stars are the examples of a higher than average mass-energy conversion efficiency in a star, which can be described mathematically so we understand its inherent brightness, its Luminosity. We know that Luminosity is simply: \[ L = \frac{energy}{time}\] and Flux is: \[ F = \frac{L}{4\pi d^2}.\]

Now we star borrowing a bit from special relativity and the definition of an electron’s cross section. So the equation with the cross section would look like: \[ \sigma_T= 6.7 \times 10^{-25} cm^2\] \[ \sigma_T \times \frac{L}{4\pi d^2} = \frac{ energy }{area \times time} \times area\] Therefore, in this equation also lies a definition for energy, and from Einstein’s equations (just accept this premise, we can go into arriving it on another blog post) we know that \[ E = pc\] where p is the momentum associated to a massless particle, i.e. a photon. Therefore, we can take our original equation and divide the energy by c to find the momentum: \[ \sigma_T \times \frac{L}{4\pi d^2 c } = \frac{ momentum }{area \times time} \times area,\] which allows for the equation to be defined for a Force: \[ \frac{ momentum }{time} = \frac{dP}{dt} = F.\] And as we saw in a previous post, the forces must equal each other to achieve hydrostatic equilibrium, in this case between gravity and radiation.
Gravitational force would be defined as: \[F_g = \frac{GM}{r^2} (m_{protons} + m_{electrons})\] and since the mass of electrons is negligible, the equations simplifies to: \[ F_g = \frac{GM}{r^2} (m_{protons}) .\]


Now we can equal our definitions of radiative and gravitational forces, and find that:  \[\frac{GM}{r^2} (m_{protons}) = \frac{L \cdot \sigma_T }{4\pi d^2 c },\] and we can now see that most of these numbers are simply constants, and we can pull out a very meaningful expression from this: \[ L = \left(\frac{4\pi G \cdot m_p \cdot c}{\sigma_T}\right) M .\] This equation directly correlates the mass of the star to its maximum inherent luminosity, the greatest brightness it could achieve were it to efficiently convert its matter into energy. This equation further simplifies down to: \[ L = 1.4 \times 10^{38} \frac{erg}{s} \left(\frac{M}{M_\odot}\right),\] all of which was discovered by Arthur Eddington, and as such this equation is the Eddington Luminosity:   \[L_E = 1.4 \times 10^{38} \frac{erg}{s} \left(\frac{M}{M_\odot}\right).\] 

Using Big (or any) Telescopes

We are bounded in a nutshell of Infinite Space: Free Form #8: Using Big (or any) Telescopes

One of the most important skills necessary for any astronomer is the ability to use and prepare telescopes in order to view everything we have talked about the last couple of months. First off, know your different types of telescopes. There are 3 main, widely used telescopes: Refractors, Reflectors, and   Catadioptric telescopes.
The Thirty Meter Telescope
http://blogs-images.forbes.com/alexknapp/files/2015/06/top-view-of-tmt-complex-1940x1089.jpg

Each of these has its advantages, but there is a reason for each. Getting Catadioptrics out of the way, they are huge and very expensive telescopes which work with a large back mirror which pools the light into an aperture just above the focusing mirror. This is the type of telescope one would use to observe deep space objects, and as such are usually the type built in places like Mauna Kea in Hawaii, and in the Atacama Desert in Chili. However, these telescopes are almost entirely motorized and have little value for the regular star-gazer, so we shall focus on ones we can readily use.
Moving on, another key telescope type is the Newtonian Reflector, which operates very much like a Catadioptric telescope, but shifts the light from the pooling mirror into an aperture on the side of the telescope, as seen below:
A Newtonian Reflector
http://ecx.images-amazon.com/images/I/51r8pTAaiKL._SX466_.jpg
Furthermore, Reflectors are very powerful telescopes since they can pool much more light than refractors (as we shall see in a moment), and can thus see farther into the night. They are usually heavier than refractors, but this is offset by their impressive angular resolution. Reflectors are used by looking into the aperture on the side of the telescope, after the light has been pooled into the aperture.
The other kind of telescope is the Refractor, a model which has gone virtually unchanged since Galileo Galilei used them. These are very long telescopes which condense light into a lens which is then pooled by a secondary lens which we observe through. Although possibly not the best kind of telescope, it is one of the most versatile and easily usable. Refractors need to be very large to see at long distances, which is how they end up in large observatories.
The Great Refractor- CfA
https://www.cfa.harvard.edu/figs/grref.gif

An observatory for a refractor has a fairly simple construction an operation. After identifying the object you wish to observe, one closes doors which would let in harsh light into the observatory, shuts off any of these lights interior to it, and turns on red lights which hardly make a difference when observing, like a dark room used in photography. Next, the dome is opened and subsequently rotated to the desired object, on which the telescope is pointed to after placing the aperture, removing the cloth over the main lens of the refractor, and turning on the motor which turns at one revolution per day (if you happen to have one of these), which helps keep the telescope trained on the object. These are the basic skills necessary to operate a telescope, aside from knowing where to observe (preferably away from urban areas) and when (most definitely not while it is raining). Happy Observing!



A Greek Astronomer for All Seasons

We are bounded in a nutshell of Infinite Space: Free Form #7: A Greek Astronomer for All Seasons

Previously, we have been looking at various Greek myths which illustrated how these viewed the stars and themselves, reflections of the heavens. However, there would be those to go beyond the limited scope of religion and culture, and push to truly understand the stars above them. First off, we have Hipparchus, possibly the greatest astronomer of the Ancient World. One of his largest contributions was the establishment of the system we still use to classify the brightness of stars and other objects: magnitudes. As intricate and complicated as this system may be, it was the first major attempt to organize the objects observed in the night sky, with which Hipparchus was able to create the largest star catalog in the Ancient World. Besides this, Hipparchus would go on to establish the nature of equinoxes and solstices, and accurately measure the movements of the Sun and Moon, which further allowed him to understand seasons and how they occurred. Hipparchus even began establishing the basis for astronomy throughout the next couple of thousand years by attempting to measure the distance to the Moon using stellar parallax. These systems were also integral to another component of his work, the development of trigonometry.  
Hipparchus
http://www.nndb.com/people/842/000103533/hipparchus-1-sized.jpg
Another interesting astronomer of the time was Eratosthenes, the man who effectively calculated the size of the Earth, created the discipline of geography, as well as being the chief librarian of the Library of Alexandria. By moving to different places across Egypt, Eratosthenes measured the differences in the elongating of a shadow, using this to create a model of a sphere which had the curvature he had measured due to the change in shadow length. From this, he extracted a proportion and accurately found the size of the Earth. Furthermore, he would go on to attempt to measure the size of the Sun, using techniques which would be imprecise with his tools, but necessary for future astronomers to effectively find this value. He, quite literally, created Geography, the science of measuring and describing the terrain of the Earth. Sadly, most of his works were destroyed in the Burning of the Library of Alexandria, but the few texts that remain and are cited by other authors tell of a man with an unsociable pursuit of knowledge.
Eratosthenes
http://web.jccc.edu/gallery/astrotext/Bills%20Files/Astronomy%20Textbook/Chapter%203_files/img2sdf.gif
These men were but two of a great tide of ancient scientists who defined the physical world for millennia, until science began to develop in the second millennium and once again the early remarks made by these astronomers were reinterpreted and researched further than they ever could have.  



Finding what Kepler devoted his life to in a couple of minutes (benefit of being in the future involved)

We are bounded in a nutshell of Infinite Space: Worksheet # 10, Problem #3: Finding what Kepler devoted his life to in a couple of minutes (benefit of being in the future involved)

3. For a planet of mass \(m\) orbiting a star of mass \(M_\star\), at \(a\) distance a in a circular orbit, start with the Virial Theorem and derive Kepler’s Third Law of motion. Assume that \(m \ll M_\star\) Remember that since m is so small, the semimajor axis, which is formally \( a = a_p + a_\star\) reduces to \(a = a_p\) (make sure you understand why).

First off, we’ll explain the concept of the axes of an ellipse. Technically, all orbits are ellipses, but many, especially in our solar system, are contained enough so they appear to be circles. However, the fact they are ellipses is what makes the solar system make sense, for the observations Tycho Brahe and Johannes Kepler made it clear ellipses were the best descriptions of these orbits, and off that Kepler created his equations. Here, the objects we are comparing are so different in size (\(m \ll M_\star\)), that the radius of either object is negligible and we can simply focus on the distance between them: \(a = a_p\).

Now, we will derive this equation a bit differently than how Kepler did it, we now being able to use the Virial theorem to describe the relationship of kinetic and potential energy. So we start with this equation: \[ K = -\frac{1}{2} U,\] and we can redefine these with the general definition of gravitational potential energy and kinetic energy, with \(a = a_p\),  \[ \frac{1}{2} mv^2 = -\frac{1}{2} \left(- \frac{GMm}{a} \right),\] which can be simplified into:  \[ v^2 = \frac{GM}{a},\] and we can now define our speed differently since we are assuming very circular orbits, we can use the definition of velocity for going around a circle once: \[ v = \frac{2\pi a}{P},\] where P is the period (time) for the rotation to occur.


Plugging that in, we have: \[\frac{4\pi^2 a^2}{P^2} = \frac{GM}{a},\] which is simplified and we find: \[ P^2=  \frac{4\pi^2 a^3}{GM},\] which is Kepler’s equation for the Period of planetary rotation around the Sun (or other central star in a solar system).  

The Reason Stars Don’t Collapse so Easily

We are bounded in a nutshell of Infinite Space: Worksheet # 9, Problem #1: The Reason Stars Don’t Collapse so Easily

1. Consider a cylindrical differential patch of fluid located a distance r from the center of a star/planet. This patch’s height is dr and the top/bottom of the cylinder has area dA. The fluid is in hydrostatic equilibrium, so this differential patch is at rest.

(a) What is the gravitational force on the cylinder if the fluid has density \(\rho(r)\) in terms of dA, dr, and the local gravitational acceleration, \(g(r)\) ?

(b) The cylinder also feels pressure force from the fluid. If the fluid pressure is \(P(r)\), what is the force associated with the pressure in terms of P, dr, and dA?

(c) Since the cylinder is at rest, the two forces have to balance each other. Write down this force balance and take \(dr \to 0\) to obtain the equation of hydrostatic equilibrium, \[\frac{dP}{dr} = - \rho (r) g(r) \]

a. This problem focusses on the concept of hydrostatic equilibrium, the point at which the pressure exerted by the mass contained in the object balances the collapse potential due to gravity. In order to understand this equation, we must first establish what the gravitational and pressure forces are.
For gravitational forces, we can imagine an infinitesimally small cylinder at a distance r from the center of our star/planet/ object with gravity, and describe the force of gravity on it as described in Newton’s equations: \[F_g = -\frac{MMG}{r^2},\] Newton’s second law of motion also describes to us the definition of a force:  \[F= ma,\] and with knowing that what we call g is the acceleration due to gravity: \[g(r) = a,\]  we can simplify the equation down to: \[ g(r) = \frac{GM}{r^2} .\] Next, we have the following factors we need to incorporate: \[ \rho (r) ~~~ dA~~~dr,\] and if we are defining the mass of this cylinder, then from density equations we know the mass is simply the volume times density, which in the case of a cylinder is the height times the area times the density: \[ dM = \rho (r) ~dA ~dr,\] and if we define this as our mass as determined by the changes in radius, we can create another \(F=ma\) equation: \[ F_{g(r)} = - \rho (r) ~dA~dr~ g(r).\]

b. Now, we will find the pressure exerted by the matter in the object, for which we have the following variables: \[ P(r) ~~~ dr ~~~ dA ,\] and from the definition of pressure we know that pressure is just \[ P = \frac{F}{A},\] so we can rewrite this expression of force and area the following way: \[ F = A~P.\] We can also deduce how the pressure varies from one radius to another, seeing how it changes across the length of the cylinder we are placing in this pressure gradient:  \[\Delta P = (P(r) - P (r + dr) ) ,\] so now it is simply a matter of placing this into our definition of force, which is: \[ F = da~\Delta P\]


c. Finally, we reconcile these two equation to find the general expression of how pressure and gravity equalize each other. Stating our definitions of forces again: \[ F_{g(r)} = - \rho (r) ~dA~dr~ g(r)\] \[ F = da~\Delta P,\] we can now equalize them and start simplifying:  \[- \rho (r) ~dA~dr~ g(r) = da~\Delta P\] \[- \rho (r) ~dr~ g(r) = \Delta P,\] now giving the full definition of \(\Delta P\): \[- \rho (r) ~dr~ g(r) = P(r) - P (r + dr) \] \[- \rho (r) g(r) = \frac{P(r) - P (r + dr)}{dr} ,\]  and the problem establishes we should take \[ dr \to 0,\] which is another way of describing a limit (which is the base for of a derivation: \[- \rho (r) g(r) = \lim_{dr\to 0} \frac{P(r) - P (r + dr)}{dr} ,\] and ultimately, we get our result, the equation for Hydrostatic Equilibrium: \[- \rho (r) g(r) = \frac{dP}{dr}.\]

Monday, February 22, 2016

The Light Comes onto Earth in Very Interesting Ways

We are bounded in a nutshell of Infinite Space: Reading #3: The Light Comes onto Earth in Very Interesting Ways

In the reading from “Astrophysics in a Nutshell” of Chapters 2.2.1-2.2.3 & 2.3, there is a quick mention of how light coming from the center of a star diffracts several times within the layers of a star before it is released into the Universe. Diffraction, in general, refers to how light changes its angle of travel once it changes medium. This happens often when you observe a fish inside a body of water, we think it’s on one place, but actually the photon bouncing off it come out at an angle which makes us think they are somewhere else. Diffraction, combined with some other characteristics of the Earth’s magnetic field and atmosphere composition, yield some of the most amazing physical phenomena humanity has encountered on Earth.

These are the Northern Lights, the Aurora Borealis (also known as the Aurora Australis in the case of the southern hemisphere), the most brilliant light show available on Earth. An Aurora is produced when the light entering the highly magnetized poles of the Earth, and thus become energized. This highly energized light then interacts with the many layers of Earth atmosphere, energizing the atoms in the gas as the light passes and changes direction repeatedly. As gases move and light diffracts, the spectacular formations of the auroras swirl and take shape. The colors of the Aurora themselves are determined by the gases present in the atmosphere as the light passes, like Neon for red and orange

Krypton for green and gray,

and Oxygen in rare cases for a majestic Red Aurora:

The difference in color stems from how these atoms interact with the light, mainly because of how atoms receive the light from the photons, and their electrons go up an energy level. Now energized, these atoms want to go back down to their ground state, which means releasing the accumulated energy, but this time with the color analogous to the energy received described by Plank’s equation: \[ E = h\nu = h \frac{c}{\lambda}\]


These lights, revered, feared, and gazed at for eons, are prime examples of how light interacts with matter, on a scale more relatable than the color of a planet several hundred or thousand light years away. 

References:
http://www.spaceweather.com/aurora/images2011/24oct11c/Dan-Salmons1.jpg
http://openwalls.com/image/19527/aurora_borealis_4_3234x2304.jpg
http://www.smh.com.au/content/dam/images/g/h/w/3/w/1/image.related.articleLeadwide.620x349.ghwn79.png/1435129354424.jpg
http://www.athropolis.com/arctic-facts/fact-nlights-color.htm
https://i.ytimg.com/vi/KAIyVZoXv9M/maxresdefault.jpg

Ever notice the moon is bigger when it’s closer to the horizon?

We are bounded in a nutshell of Infinite Space: Free Form #6: Ever notice the moon is bigger when it’s closer to the horizon?

Has the Moon ever looked bigger? Does it change its distance to the Earth throughout the night? Could the Earth’s atmosphere or the amount of photons reflected off the moon have anything to do with it?

Actually, none of these reasons explain why the Moon is perceived by most to be bigger at certain parts of the night, rather the answer actually comes down to an optical illusion. The moon, when close to the horizon, is also much closer to the objects near the ground, giving the human eye a comparison from which to estimate the Moon’s size, but this actually leads us astray. Records since the Ancient Greeks to Immanuel Kant speak as to why this illusion occurs, but Aristotle was quite wrong when it came to Astronomy, as you could tell if you ever read his dialogues on falling objects. He postulated that the Earth’s atmosphere created a lensing effect near the horizon, while there was no evidence if this happening with any other object besides the Sun.

The optical illusion occurs from the eye’s estimate of size based on depth perception, which noticeably changes when the Moon is closer to other objects of similar or apparently larger size (i.e. trees, mountains, among others). Otherwise, when the Moon is well above the horizon, the only point of comparison is the vastness of space, which most definitely makes our eyes believe we are looking at something very far away. Actually, when close to the horizon, the Moon is farther away by a distance slightly longer than the radius of the Earth simply because of how the Earth’s rotation changes our frame of reference. This effect is called the Ebbinghaus Illusion (popularized in a psychology textbook published in 1901), which places two objects of the same size in the middle of other objects of smaller and larger size.


As can be seen in the image, the object placed in between large objects is perceptibly smaller than the object between smaller objects. This is the same effect for the moon, where large expanses of empty space make the Moon seem small (its ordinary size) and when close to the horizon, the smaller trees and relief make the Moon seem like another planet close to ours. In fact, if you were to simply close one of your eyes as you observed the Moon when it’s close to the horizon, you would see it shrink to its regular size. The same goes if you were to block out the features of the horizon and were only looking at the moon. 

References:
https://en.wikipedia.org/wiki/Moon_illusion
https://upload.wikimedia.org/wikipedia/commons/thumb/b/bc/Mond-vergleich.svg/2000px-Mond-vergleich.svg.png
http://i.livescience.com/images/i/000/049/591/original/super-moon-2011-tim-mccord-entiat-wash.JPG?1336008254

View the Stars like a Greek!

We are bounded in a nutshell of Infinite Space: Free Form #5: View the Stars like a Greek!

The Greek civilization is perhaps the best to refer to when describing ancient Astronomy, in part because of the incredible amount of records which have been kept across millennia. Aside from the wonderful stories which describe how certain stellar constellations got their names, there are a couple of stories which describe some of the basic tenets which were held for centuries and millennia. This is the case of the origin for the word Galaxy.

As the myth goes, Zeus, the king of the Greek gods, ruler of the skies and Olympus, would impregnate mortal women, and his children through them were the greatest heroes ever to walk the face of the Earth. One such case was Heracles (‘Ηρακλεως), the son of Zeus and Alcmene, who was gifted with god-like strength, endurance, and hubris. However, when Heracles was being raised, Zeus wished he would be yearned with the mother’s milk of a goddess, specifically: Hera. Aside, Heracles is an interesting word play in ancient Greek, since his name literally means, the Glory of Hera, but if you remember anything about their stories, you know they mutually hated each other, to the point Hera caused the incident which led to Heracles’ 12 labors, as well as turning Alcmene into a cow. Anyways, Zeus attempted to have Heracles suckle Hera when we was a baby, placing him on her while she was asleep. As the infant gained the powers of a god, Hera would wake and swat him aside, but her milk would spill from her home in the skies. This spill would be the group of stars concentrated around a patch in the sky, which was the Milky Way, both figuratively and literally.  After all, the word Galaxy comes from the Greek Γαλαχτος, which literally means “of milk”.
Jacopo Tintoretto -The Origin of the Milky Way
from https://upload.wikimedia.org/wikipedia/commons/8/80/Jacopo_Tintoretto_-_The_Origin_of_the_Milky_Way_-_Yorck_Project.jpg

One more story for this time: Orion, the hunter. Orion was one of the best hunters in all of Greece, so great that he rivaled one of the gods with his prowess, Artemis, the goddess of the hunt, maidens, and the moon. Orion and she would go on to become great friends, hunting the greatest beasts in mythical lore, so much so Apollo, Artemis’s brother, became jealous. Apollo would send a giant scorpion to lead Apollo to the sea, while he convinced Artemis that Orion had raped her priestesses. Artemis would seek out and kill Orion as he swam into a lagoon, but later finding her priestess all right and the treachery unmasked, she placed the mighty hunter among the stars. This figure, the three stars Alnitak, Alnilam and Mintaka making up its belt, is one of the most prominent groupings of stars in the night sky, co much so every major civilization we know of built temples to represent them (i.e. the Great Pyramids of Giza), and have remained a staple of the coming of winter, and for many Hispanic cultures, the arrival of the Three Wise Men to bring gifts to children who remember their story.

There are millennia’s worth of stories written on the stars, one only needs a little searching to hear them again, as they once were on the isle of Lesbos.  

References: 
https://upload.wikimedia.org/wikipedia/commons/8/80/Jacopo_Tintoretto_-_The_Origin_of_the_Milky_Way_-_Yorck_Project.jpg
http://www.comfychair.org/~cmbell/myth/myth.html
http://people.ucsc.edu/~rosewood/writing/tales/orion.htm
http://www.ancient-code.com/wp-content/uploads/2015/12/76c29uIpyz-min.jpg
http://www.huffingtonpost.com/mario-livio/our-home-galaxy-myths-and-facts_b_1914174.html

Thought the universe was empty? Guess again, it’s actually emptier.

We are bounded in a nutshell of Infinite Space: Worksheet # 8, Problem #1: Thought the universe was empty? Guess again, it’s actually emptier.

1. The matter between stars is known as the interstellar medium and is composed of mostly dust and gas. This dust absorbs stellar light and therefore reduces the apparent brightness of stars as viewed from the Earth. This effect is known as extinction and worsens as we observe more distant stars.

(a) Using trigonometric parallax you measure that the distance to a particular star is 200pc. You know that the absolute magnitude of this star should be 2 and measure an apparent magnitude of 12. How much flux have you lost due to the intervening dust? Compute your answer in magnitudes.

(b) Optical depth is a measure of the absorption of photons as they travel through a medium. The definition of optical depth is: \[ \tau = \ln(I_{in} / I_{out})\] where I denotes the specific intensity of the source. In this problem, we are looking at a single source at a fixed distance, so we can also express \(\tau\) in terms of the flux: \[ \tau = \ln(F_{in} / F_{out}) \] Use this latter definition to determine the relationship between \(\tau\) and apparent magnitude. What is the optical depth along the line of sight to the star in the previous problem?

(c) We can now compute the amount of dust required along our line of sight to produce the observed extinction. The optical depth along a line of sight can be calculated from the absorption cross section of any intervening material and the number density of the particles with that cross section. This can be written \[ \tau = N \times \sigma\] , where N is the total number of particles per \(cm^{-2}\) in the line of sight and \(\sigma\) is the cross section of the individual particles.

Assume that each dust grain has the typical size of \(r = 0.1 \mu m \) and is spherical. Calculate the geometric cross section of a dust grain in units of \(cm^2\). Assume that this geometric cross section is the absorption cross section, which is true for visible light. Determine how many particles per \(cm^2 \) you would need to obtain the calculated optical depth. This value is referred to as a “column number density.”

(d) We know that the mass in gas is about 100 times more than in dust. What is the column number density of gas along the same line of sight?

(e) What is the average gas density along the same line of sight? Express your answer as the number of particles per \(cm^3\) .

a. Dust is part of the interstellar medium, present in (very) high orders of magnitude, dust affects most astronomical observations, and can lead to drastic changes of the ideal observations. This is the case for the object we are observing which has an apparent magnitude of 12, but according to its abosolute magnitude of 2, we can solve and find that: \[M = m - 5 (\log_{10} (d) - 1)\] \[m = M + 5 (\log_{10} (d) - 1)\]  \[m = 2 + 5 (\log_{10} (200) - 1)\] \[m = 2+ 5 (1.5),\] the apparent magnitude we should be receiving is: \[ m_{calculated} = 8.5,\] and we know that: \[ m_{measured} = 12 ,\] so the \[ m_{discrepancy} = ~Loss~ of ~Flux~ = 3.5 \]

b. Another important component is the relationship of intensity and flux describes as: \[ \tau = \ln(F_{in} / F_{out}) ,\] and we also know that  \[ \frac{F}{F_0} = 10^{-0.4(m - m_0)}, \] as per the basic equations relating magnitudes we have been using thus far. And this is analogous to \[ F_{in} / F_{out} = F~ /~ F_0,\] so \[ \tau = \ln \left(10^{-0.4(m - m_0)} \right) ,\] and in the case of the apparent magnitudes we had calculated and observed, we find that: \[ \tau = \ln \left(10^{-0.4(8.5 - 12)} \right) ,\] \[ \tau = 3.22 ,\] which is the so-called optical depth.

c. Furthermore, from the equation given in the problem, we can see how: \[ \tau = N \times \sigma, \] which describes the optical density as the number of particles in a cross section of the sky. Furthermore, the cross section is simply:  \[ \sigma = \pi r^2 ,\] where the radius is that of a dust particle, this being:  \[ \sigma = \pi (0.1 \mu m )^2 \] \[ \sigma = \pi (10^{-5} cm)^2\] \[ \sigma = \pi 10^{-10} cm^2.\] now we can plug these numbers to find the Column Number Density where the optical depth is what we calculated in the precious problem:  \[ N = \frac{\tau}{\sigma}\] \[ N = \frac{3.22}{\pi 10^{-10} cm^2}\] \[N = 1.02 \times 10^{10} cm^{-2},\] which is the amount of dust particles per optical column (200 pc long in this case) .  

d. Next, we can find the amount of atoms considering their collective mass is 100 times that of the collective of dust particles. To find the dust particles mass, we multiply their volume, density, and number to get the total dust in a column: \[ V_{dust} \cdot N_{dust} \cdot \rho_{dust} = M_{dust/cm^2},\] where the density is the density of most silicates:  \[\rho_{dust} = 3 \frac{g}{cm^3},\]  and the volume is simply: \[ V = \frac{4}{3} \pi r^3 \] \[ V_ dust = \frac{4}{3} \pi (10^{-5} cm)^2 = 4 \times 10^{-15} cm^3,\] and we know the column number density from the last problem: \[N = 1.02 \times 10^{10} cm^{-2}.\] Plugging in, we get: \[M_{dust/cm^2} = V_{dust} \cdot N_{dust} \cdot \rho_{dust} \] \[ M_{dust/cm^2} =  4 \times 10^{-15} cm^3 \cdot 1.02 \times 10^{10} cm^{-2} \cdot 3 \frac{g}{cm^3} \] which tells us the mass of all the dust in the column is: \[ M_{dust/cm^2} =  1.2 \times 10^{-4} \frac{g}{cm^{2}}\]

Now we just multiply this mass by 100 and find the mass of all the gas ( which we can assume is predominantly Hydrogen):  \[ M_{gas/cm^2} = 100 M_{dust/cm^2} \] \[ M_{gas/cm^2}= 100 \cdot 1.2 \times 10^{-4} \frac{g}{cm^{2}}\] \[ M_{gas/cm^2} =  1.2 \times 10^{-2} \frac{g}{cm^{2}}\] and now we can just plug in the mass of a hydrogen atom to find the amount of atoms: \[M_{H ~atoms} = 1.67 \times 10^{-24} \frac{g}{atom},\] so we have: \[ \frac{ M_{gas/cm^2} }{ M_{H ~atom} } = Particles / cm^2\] \[ \frac{ M_{gas/cm^2} }{ M_{H ~atom} } = \frac{ 1.2 \times 10^{-2} \frac{g}{cm^{2}}}{1.67 \times 10^{-24} \frac{g}{atom}},\] which is: \[ Particles / cm^2 = 7.2 \times 10^{21} \frac{atoms}{cm^2},\] in the entire column we are observing.  

e. But, all these calculations have been on the basis that we are talking about the entire column, what about a simple \(cm^3\) ? We know the column is 200 pc long, and this equates to:  \[ 1 pc = 3.1 \times 10^{18} cm\] and there are\[ d = 200 pc,\] therefore: \[ Particles/ cm^3 = \frac{ H ~atoms / cm^2 }{d} \] so it is just a matter of plugging in and we find: \[ Particles / cm^3 = \frac{7.2 \times 10^{21} \frac{atoms}{cm^2}}{200 \times 3.1 \times 10^{18} cm }\]  \[ Particles / cm^3 = 11.6 \frac{atoms}{cm^3},\] which serves to show us how little there actually is in the vastness of space, rather we only see it once we start looking at the objects in incredibly large perspectives.


Judging a Star’s Actual Brightness

We are bounded in a nutshell of Infinite Space: Worksheet # 7, Problem #2: Judging a Star’s actual Brightness

2. Okay, let’s use your new-found knowledge of magnitudes.

(a) Suppose you are observing two stars, Star A and Star B. Star A is 3 magnitudes fainter than Star B. How much longer do you need to observe Star A to collect the same amount of energy in your detector as you do for Star B?

(b) Stars have both an apparent magnitude, m, which is how bright they appear from the Earth. They also have an absolute magnitude, M, which is the apparent magnitude a star would have at \( d = 10 ~parsec\) (One parsec = 3.26 light years. We’ll learn more about this quantity in the next question). How does the apparent magnitude, m, of a star with absolute magnitude M, depend on its distance, d away from you?

(c) Assume that the part of the Galaxy that we live in is uniformly filled with stars with a number density \(n_\star \). Suppose you plan to measure the spectrum of every star down to an apparent magnitude m. How many more stars do you have to observe if you observe to a limit that is one magnitude fainter? In other words, what is the fractional change (to one digit) in the number of targets for each magnitude fainter you can observe? (Hint: Don’t overthink this problem. We don’t have enough information about the distribution of stars to consider specifics.)

a. Ok, so we’ve worked with magnitudes before, an ancient system of measuring a stellar object’s brightness. But in this case we have to prove the equations we have been using thus far. First off, let us remember how magnitudes work: the higher the number, the fainter the star is, and this scale works in a logarithmic relationship where each increase in magnitude corresponds to an increase in flux as according to the equation:  \[ \frac{F_1}{F_2} = 10^{0.4(m_2 - m_1)}.\] Now then, we can use this equation to work through the current problem we are tasked with.

We know from the problem that \[ m_A = m_B - 3\] \[\Delta m = 3,\] and so this difference in magnitude corresponds to a “precise” (not the best word considering most of these systems are approximations half the time) flux ratio, as seen in the equation above. Furthermore, we can rewrite the same equation in the following way: \[ \frac{F_1}{F_2} = 10^{0.4(m_2 - m_1)}\] \[ m_1 - m_2 = -2.5 \log_{10} \left(\frac{F_1}{F_2} \right) ,\] by changing the way we use the logarithmic bases. Now, we just plug in the right definitions within the context of this problem and get: \[ m_ A - m_B = -2.5 \log_{10} \left(\frac{F_A}{F_B} \right) ,\] and now plugging in real numbers: \[ m_ B - 3 - m_B = -2.5 \log_{10} \left(\frac{F_A}{F_B} \right) \]  \[ -3 = -2.5 \log_{10} \left(\frac{F_A}{F_B} \right) \] \[\frac{6}{5} = \log_{10} \left(\frac{F_A}{F_B} \right) ,\]  which finally yields: \[ \frac{F_A}{F_B} = 10^(6/5) = 15.8,\] which describes the ratio of time necessary to collect the same amount of light for each star, meaning Star A takes 15.8 times longer than Star B.

b. First, we shall define in a broader sense what the Flux ratio is. From the simple definition of Flux we know it relates surface area and Luminosity, which can be simplified once we assume several things are the same, which is what we do here: \[ \frac{F_A}{F_a} = \frac{\frac{L_A}{4\pi d_A ^2}}{\frac{L_a}{4\pi d_a ^2}} = \frac{d_a ^2}{d_A ^2 }, \] and now, knowing this star has an Absolute Magnitude M, which we know is the brightness at 10 pc, therefore we can establish what is the relationship of a star’s absolute and apparent magnitude because of its distance. So if the Absolute magnitude is at 10 pc: \[\frac{d_a ^2}{d_A ^2 } = \frac{d_a ^2}{ 100 ~pc^2}, \] which we will plug back into the original equation.  

Knowing the equation from the last problem, we change it so it represents the star we are looking at now: \[ m_1 - m_2 = -2.5 \log_{10} \left(\frac{F_1}{F_2} \right) \]  \[ m_A - m_a = -2.5 \log_{10} \left(\frac{F_A}{F_a} \right), \] which is also just a way of describing absolute and apparent magnitudes (we now plug in our definition of distance in this case): \[ M - m = -2.5 \log_{10} \left(\frac{d_a ^2}{ 100} \right) ,\] and we can start simplifying down with the properties of logarithms:  \[ M - m = -2.5 \log_{10} (d_a ^2) - \log_{10} (100) \]  \[ M - m = -2.5 (2 \log_{10} (d) - 2) ,\] which yields the equation describing the relationship of apparent and absolute magnitude, also known as a use for the distance modulus: \[ M = m - 5 \log (d) - 5.\]

c. In this problem, we simply want to understand how many more stars, of they were uniformly arranged, we would see if we looked one magnitude fainter than we previously did. Therefore, we start setting up with the standard equation for relating two  magnitudes: \[ m_x - m_{x_0} = -2.5 \log_{10} \left(\frac{F_x}{F_{x_0} } \right) ,\]  and we know the new stars we want to see are of 1 magnitude greater than our base stars: \[ 1  = -2.5 \log_{10} \left(\frac{F_x}{F_{x_0} } \right) ,\] Now just a matter of simplifying for a bit (and adding the broader definition of flux we saw earlier since the stars are uniform), we see that: \[ -\frac{2}{5} = \log_{10} \left(\frac{\frac{L_x}{4\pi d_x ^2}}{\frac{L_{x_0} }{4\pi d_{x_0}  ^2}}  \right) \] \[ -\frac{2}{5} = \log_{10} \left(\frac{d_{x_0}  ^2 } {d_x ^2}\right)\] \[ \frac{d_{x_0} ^2 }{ d_x ^2 } = 10^{-0.4},\] which means we can find a direct relationship between the distances we can see out two now by increasing our observation to one more magnitude: \[ \frac{d_{x_0} }{ d_x} = 0.63 \] \[ d_{x_0} = 0.63 ~d_x.\]

This increased radius of observation can then be used in a density equation where we substituted out the mass for number of stars: \[N_\star = V \times n_\star\] \[ \frac{N_{\star~x}}{N_{\star~ {x_0}}} =\frac{V_{x} \times n_\star}{V_{x_0}  \times n_\star},\] where we define the spherical volume by: \[ V = \frac{4}{3} ~\pi ~r^3,\] and placing it in the equation we get: \[ \frac{N_{\star~x}}{N_{\star~ {x_0}}} = \frac{\frac{4}{3} ~\pi ~d_x ^3 \times n_\star}{\frac{4}{3} ~\pi ~d_{x_0}^3 \times n_\star},\] which can be simplified down to: \[ \frac{N_{\star~x}}{N_{\star~ {x_0}}} = \frac{d_x ^3}{ d_{x_0} ^3},\] which in turn can be plugged in with the extension of the radius of observation we calculated earlier: \[ \frac{N_{\star~x}}{N_{\star~ {x_0}}} = \frac{d_x ^3}{ (0.63~ d_x) ^3} \] \[ \frac{N_{\star~x}}{N_{\star~ {x_0}}} = \frac{1}{0.25},\] which means by gaining a magnitude of observation sensitivity, we increase the amount of stars we can see by a factor of: \[ \frac{N_{\star~x}}{N_{\star~ {x_0}}} = 4.\]



Monday, February 15, 2016

Why aren’t stars perfect Blackbodies?

We are bounded in a nutshell of Infinite Space: Reading #2: Why aren’t stars perfect Blackbodies?
All this time, we have been explaining how stars are easily approximated as blackbodies, which offers us great insights into their properties. However, there is a condition which limits the possibility of establishing a sure connection between the theoretical approximation of a blackbody and a huge mass of nuclear fusion held together and pulled apart by gravity and pressure. These are the dips in a spectral image of a star:
So what is the source of these dips, for which we travel to the early nineteenth century, where spectral lines are about to be discovered. In 1802, William Wollaston was the first we know to have seen and describe the dark lines which appeared after splitting the light stemming from the Sun with a prism. Later on, Joseph Von Fraunhofer, a German optician, successfully split the spectral lines of the Sun and catalogued all 475 according to their width and position. These spectral lines are now known as Fraunhofer lines, and they were used to find the elements which compose the Sun.

While observing these lines, Fraunhofer was able to pick out the specific lines which make up the specific color of salt when thrown into a flame. Using this same technique, Robert Bunsen would create a burner which most precisely yielded the color spectrum of different elements when exposed to the flame. Along with Gustav Kirchhoff, Bunsen found how these specific lines in the Fraunhofer solar spectrum corresponded directly with the lines emitted by other elements, as is the case of iron and what would later be determined as Helium in 1868. Kirchhoff, being a theoretical physicist, established a set of descriptions to broadly explained spectral lines, the most relevant being how cool, diffuse gas could absorb energy from the full spectrum and create absorption lines.


These absorption lines also helped other scientists and astronomers find a clear relationship between slightly different spectra, caused by what we then learned was the phase shift from the Doppler effect of moving while emitting photons. This equation is described as: \[\frac{\lambda_{obs} - \lambda_{rest}}{\lambda_{rest}} = \frac{v_r}{c},\] where \(v_r\) is the radial velocity, c is the speed of light, and the wavelengths are differentiated form being the observed ones and the normally emitted. These early descriptions of moving objects would be the basis of Hubble’s expanding universe model, based on a shift in the wavelength of atomic hydrogen from the established standard, z, also known as redshift, is calculated precisely the same as the radial velocity is. Spectral lines are the basis for our understanding of what other objects are made of, and are what has allowed us to better see how other systems develop, accumulate matter, and how stars forge the heaviest elements in spectacular explosions we call supernovae. 

References: 
Maoz, D. Astrophysics in a Nutshell.
Carrol & Ostlie; 2007; Introduction to Modern Astrophysics
https://upload.wikimedia.org/wikipedia/commons/thumb/2/2f/Fraunhofer_lines.svg/2000px-Fraunhofer_lines.svg.png

Egyptian Astronomy

We are bounded in a nutshell of Infinite Space: Free Form #4: Egyptian Astronomy


One of the oldest human civilizations, the Ancient Egyptians made some of the greatest strides in the advancement of knowledge in the ancient world. Between developing a highly sophisticated writing structure to creating some of the greatest feats of construction in human history, these peoples ordered the world in their particular way, attributing deities and divine interventions to describe the commonplace events of every day. However, there was an incredible development of mathematics by the Egyptians which allowed them to find and describe much more than simply a Sun god racing across the sky every day. They understood that the day was approximately 365 days long, and so developed a calendar with 12 months, 30 days to each, and 5 extra ceremonial days. They also understood that every day was 24 hours long, and so divided the hours so as to be sure there would always be 12 hours of night and 12 of day. 

For the Egyptians, these numbers also had religious significance, as is the case of the hours in a day: these represent how Ra, the Sun, born from Mehet-Weret, the Sky (in ancient traditions), rode across the sky every day in his fiery chariot, to then sink into the Underworld every night, sailing across the Duat into the domain of Osiris, the god of the Underworld. During Ra’s battles at night, the Egyptians would see the stars, represented by the goddess of writing, Seshat, and the Milky Way itself was the birth of Ra by Mehet-Weret, while the moon was represented by the god of wisdom, Thoth. To think, the Egyptians were so transfixed by the stars that they built monuments to honor and mirror them. This is the case of the Great Pyramids in Giza, each of them corresponding to Alnitak, Alnilam, and Mintaka (as we know them now), the three stars that make up Orion’s belt, or rather Osiris’s, according to Egyptian star charts.

Here is yet another interesting story, how when Osiris and his brother Set fought for control of Ancient Egypt, Set would finally defeat Osiris destroy his body and scatter it everywhere, including the constellation which would be named after him. Isis, Osiris’s sister and wife, would collect Osiris’s body and reconstruct him to bear their child, Horus, who would eventually defeat Set. Now, Osiris is never seen during the day, for that is Ra’s domain, but at night, when Ra crosses the Duat, Osiris is seen clear above the largest tombs ever constructed, the pyramids. Furthermore, there is an important aspect to the year and Egyptian Astronomy and religion which further explains the origin of the 5 extra days to complete the calendar. Nut, who inherits being goddess of the stars, moon, night, sky, astronomy, and the cosmos, wished to have children with her husband, Geb, the Earth, but Ra decreed she could not have them on any day of the year. Therefore, with Thoth’s help, Nut stole moonlight from its goddess, Khonsu, and used it to create 5 extra days in the year, thus allowing her to have her 5 children: Osiris, Isis, Set, Nephthys, and Horus (in some myths). 


Overall, the Egyptians had a deep reverence towards the celestial sphere, seeing in it their origins and their future, literally and figuratively. They recognized how the world was much more than themselves, and would later give rise to the most prolific ancient civilization, the Alexandrian scholars. 

References: 
http://www.pifeed.net/wp-content/uploads/2015/04/Great-Pyramids-of-Giza.jpg
https://upload.wikimedia.org/wikipedia/commons/f/f3/NutMAM.jpg
http://egyptianmythology.org/wp-content/uploads/2015/05/Nun-Raises-the-Sun-BC.jpg
http://www.starteachastronomy.com/egyptian.html
https://explorable.com/egyptian-astronomy
https://en.wikipedia.org/wiki/Egyptian_astronomy

Gravitational Waves, Now what do we do with them?

We are bounded in a nutshell of Infinite Space: Free Form #3: Gravitational Waves, Now what do we do with them?
“On September 14, 2015 at 09:50:45 UTC the two detectors of the Laser Interferometer Gravitational-Wave Observatory simultaneously observed a transient gravitational-wave signal. The signal sweeps upwards in frequency from 35 to 250 Hz with a peak gravitational-wave strain of \(1.0 \times 10^{-21}\). It matches the waveform predicted by general relativity for the inspiral and merger of a pair of black holes and the ringdown of the resulting single black hole. The signal was observed with a matched-filter signal-to-noise ratio of 24 and a false alarm rate estimated to be less than 1 event per 203 000 years, equivalent to a significance greater than \(5.1 \sigma\). The source lies at a luminosity distance of \(410^{+160}_{-180}\) Mpc corresponding to a redshift \( z = 0.09^{+0.03}_{-0.04}\) . In the source frame, the initial black hole masses are \(36^{+5}_{-4} M_\odot\) and \(29^{+4}_{-4} M_\odot\), and the final black hole mass is \( 62^{+4}_{-4} M_\odot\), with \(3.0^{+0.5}_{-0.5} M_\odot c^2\) radiated in gravitational waves. All uncertainties define 90% credible intervals. These observations demonstrate the existence of binary stellar-mass black hole systems. This is the first direct detection of gravitational waves and the first observation of a binary black hole merger.”

This is the abstract for the announcement last week for the first direct observation of Gravitational Waves, further confirming Einstein’s Theory of General Relativity, and pushing us to a new frontier of Astrophysics and Observation. We previously discussed the history of Gravitational Waves (see http://ay16-rodrigocordova.blogspot.com/2016/02/gravitational-waves-next-e-m-spectrum.html) and how they could potentially revolutionize astrophysical observation, well now this is a reality! As the abstract describes, the observed gravitational waves had a significance of \(5.1 \sigma\), which means there is great certainty that the observation is correct and real. This significance stems from the amount of falsifiability of the data, meaning the lack of it being anything other than the proposed explanation, two black holes circling and colliding with one another. Furthermore, the authors go on to explain in the later parts of the paper (found here: https://journals.aps.org/prl/pdf/10.1103/PhysRevLett.116.061102) how the waves were confirmed to move at the speed of light by knowing the distance between the two LIGO detectors (one in Louisiana and one in Washington State) and measuring how the detection lagged one behind the other by a time equivalent to their separation distance divided by the speed of light. This, along with other factors, gives us one of the most important aspects of the discovery: the fabric of space-time is real, and it allows for the propagation of gravitational forces at the speed of light. However, I do not see these as the most relevant consequence of the discovery, rather the prospects it offers are much more enthralling.

The most significant portion of this discovery is the opening of an entire new realm of fields in observational astronomy, and Astrophysics. Now we have the basis to peer deeper into the universe, into its hidden dimensions and a new range of possible objects to be studied and to be gained insight from. From the paper, we know they detected several other sources and objects that were detected, although they were not precise enough to be catalogued as a verifiable event. This hints at how common gravitational wave events could be, sufficiently so that we could see completely new facets of the universe’s construction we haven’t even dreamt of yet.  Within the coming years, Astrophysics will change to accommodate what is the first real shift in observational astronomy since the Ancients, we now have a completely different spectrum to work with, not based on Electromagnetism, but rather the interactions of large masses and the possible reasons for that (we’ll discuss gravitons on another occasion). Regardless of what you may think of physics and astrophysics, how we may just be getting close to explaining mos of the phenomenon we see, events like there are what lead us to other discoveries and new thing to describe and attempt to understand, because for all we know, general relativity might just be a special case of some other, more general, explanation of the universe(s).

References: 

Taking the Sun’s Temperature

We are bounded in a nutshell of Infinite Space: Worksheet # 6, Problem #2: Taking the Sun’s Temperature

2. Consider the amount of energy produced by the Sun per unit time, also known as the bolometric luminosity, \(L_{\odot}\) . The same amount of energy per time is present at the surface of all spheres centered on the Sun at distance \(r > R_{\odot}\) . However, the flux at a given patch on the surface of these spheres depends on r. (the Stefan-Boltzmann Constant: \(\sigma = 5.7 \times 10^{-5} erg~ cm^{-2}~ s^{-1} ~ K^{-4} \) )
               
            a. How does flux, F, depend on luminosity, L, and distance, r?
               
            b. The solar flux at the Earth-Sun distance has been measured to high precision, and for the purposes of this exercise is given by \(F_{\oplus} = 1.4 \times 10^{6} ergs~s^{-1}~cm^{-2} \) . Given that the Sun’s angular diameter is \( \theta = 0.57 \) degrees, what is the effective temperature of the Sun? (Hint: start with the mathematical version of the first sentence of this problem, namely: \[L_\odot (r = R_\odot ) = L_\odot (r = 1 AU)\] and then expand the left and tight sides of their respective distances and fluxes at those distances. )

a. From the several problems we have already worked on, as well as a simple analysis of the units of flux and luminosity, we know they are related by a factor of the area covered. We also know celestial sources emit light isotopically (in all directions) and their light seems dimmer the farther they are away. Therefore, it is just an understanding that flux (the brightness we see) is lessened by the distance (and area) covered by the photons is best described as:  \[F = \frac{L}{4\pi r^2} \]

b. Knowing the base values described in the problem, we can start probing into the actual uses of the equations which describe blackbodies, and objects that are very close to them (i.e. stars). One of the first things we will need before using the relationships of Flux, Luminosities and Radii, we need to solve a bit of trigonometry.

As you can see in the image, and in the problem’s descriptions, the angular size of the Sun in our sky is 0.57 degrees. It’s being so small means we can use a small angle approximation to find its diameter, the opposite side to a right triangle, where the adjacent side is the distance from the Earth to the Sun, So, we have: \[\theta = \frac{D_{\odot}}{d_{\odot - \oplus}}\] and the angle needs to be used in radians: \[\theta = 0.57 degrees \times \frac{\pi ~rad}{ 180 degrees} = 0.01 ~rad\] Now we can solve for the diameter of the Sun: \[ \theta = \frac{D_{\odot } }{1 AU}\] \[ 0.01 =\frac{D_{\odot}} {1 AU} \] \[ 0.01 AU = D_\odot ,\] and with a simple division by 2: \[ 0.005 AU = R_\odot \]

Now we get into the more interesting part. Using equations which describe blackbodies, we know Luminosity is intrinsically related to Temperature. But first, we use the equations relating luminosity and flux. From the problem, we have: \[L_\odot (r = R_\odot ) = L_\odot (r = 1 AU), \]  and \[F = \frac{L}{4\pi r^2} ,\] and these two can be combined as: \[ F_{R_\odot} (4\pi R_\odot ^2 ) = F_{ d_{\odot - \oplus}} (4 \pi d_{\odot - \oplus} ^2 ),\] and after cancelling out some constants:  \[ F_{R_\odot} (R_\odot ^2 ) = F_{ d_{\odot - \oplus}} (d_{\odot - \oplus} ^2 ),\] we can start plugging in \[ F_{R_\odot} (0.005~AU)^2  = F_{ d_{\odot - \oplus}} (1~AU)^2 ,\] and with some facts retrieved from the problem statement: \[ F_{R_\odot} (0.000025) =  1.4 \times 10^{6} ergs~s^{-1}~cm^{-2} ,\] and the Flux at the surface of the Sun is: \[ F_{R_\odot} = 5.6 \times 10^{10}  ergs~s^{-1}~cm^{-2}.\]


Now, we employ the equation derived in a previous blog post (), which gives a clear relationship between Luminosity and Temperature, but in this case we will be replacing it with the Flux at the surface of the Sun: \[ L = 4\pi R^2  ~\sigma T^4 ,\] and changing it into Flux: \[ F_{R_\odot} \cdot 4\pi R_{\odot}^2 = 4\pi R_{\odot}^2 ~ \sigma T^4 ,\] now we cancel some constants and begin to plug in: \[ F_{R_\odot} = \sigma T^4\] \[ T = \left( \frac{ F_{R_\odot} }{\sigma }\right)^{\frac{1}{4}} \] \[ T =\left( \frac{ 5.6 \times 10^{10} ergs~s^{-1}~cm^{-2} }{ 5.7 \times 10^{-5} erg~ cm^{-2}~ s^{-1} ~ K^{-4}}\right)^{\frac{1}{4}} ,\] and the final solution, with a bit of approximation, is: \[ T_\odot = 5600 K \] 

Let’s get into the math of Stellar Temperatures

We are bounded in a nutshell of Infinite Space: Worksheet # 5, Problem #2: Let’s get into the math of Stellar Temperatures

2. Blackbodies are nice because they’re such simple objects. Their outward appearance is entirely determined by their temperature. If there were cows in space, astronomers would imagine them to be spherical blackbodies (and seriously, it wouldn’t be a bad approximation). In this exercise we’ll take advantage of the relative simplicity of blackbodies to derive some useful expressions that you’ll use this term, and throughout tour astronomy career.

a. In astronomy, it is often useful to deal with something called the “bolometric flux”, ot the energy per area per time, independent of frequency. Integrate the blackbody flux \(F_\nu (T)\) over all the frequencies to obtain the bolometric flux emitted from a blackbody, \(F (T) \). You can do this using u substitution for the variable \(u \equiv hv/kT\) . This will allow you to split thing s into a temperature dependent term, and a term comprising an integral over all frequencies. However, rather than solving for the integral, just set all the constants to a new, single constant called \(\sigma\) , which is also known as the Stefan-Boltzmann constant, and retain the dependence on the temperature T. If you’re really into calculus, go ahead and show that \(\sigma \approx 5.7 \times 10^{-5} erg~s^{-1}~cm^{-2}~K^{-4}\).  Otherwise, commit this number to memory.

b. The Wein Displacement Law: Convert the units of the blackbody intensity from \( B_\nu (T) \) to \(B_\lambda (T) \) IMPORTANT: Remember that the amount of energy in a frequency interval \(d\nu\) has to be exactly equal to the amount of energy in the corresponding wavelength interval \(d\lambda\).

c. Derive the expression for the wavelength \(\lambda_{max}\) corresponding to the peak of the intensity distribution at a given temperature \(T\). (HINT: How do you find the maximum of a function? Once you do this, again substitute \(u \equiv h\nu / kT\) ). The expression you end up with will be transcendental, but you can solve it easily to the first order, which is good enough for this exercise.

d. The Rayleigh-Jeans Tail: Next, let’s consider photon energies that are much smaller than the thermal energy. Use a first-order Taylor expansion on the term \(e^{\frac{h\nu}{kT}}\) to derive a simplified form of \(B_\nu (T)\) in this low-energy regime. (HINT: The Taylor Expansion of \(e^x \approx 1 + x\) .)

e. Write an expression for the total power output of a blackbody with a radius R, starting with the expression for \(F_\nu\). This total energy output per unit time is also known as the bolometric luminosity, L.

a) As we have seen in less detail at other points, blackbodies are a good approximation for most objects in the universe. The most “perfect” blackbody we have found is actually the CMB (Cosmic Microwave Background), the afterglow of the Big Bang. Based on a relationship between the frequency/wavelength of the observed light and the temperature of the light, we find an equation which can relate these two, as well as be derived and anti-derived to give broader descriptions.  

From a previous problem, we know that a description of the Flux of a blackbody, any object which absorbs and releases thermal radiation, is represented by: \[ F_\nu (T) = \frac{2\pi}{c^2} \frac{\nu^3 h}{e^{h\nu/kT} -1} ,\] so if we integrate over all possible frequencies, going to infinity, we have: \[ \int _0 ^\infty F_\nu (T) = \frac{2\pi h }{c^2} \int _0 ^\infty \frac{\nu^3 }{e^{h\nu/kT} -1} d\nu ,\] where we can substitute \[ u = \frac{h\nu}{kT} \] and \[u\frac{kT}{h} = \nu\] as well as \[du = \frac{h}{kT}, \] into \[ F(T) = \int _0 ^\infty \frac{\left(\frac{kT}{h}\right) u^3 }{e^{u} -1} \left(\frac{kT}{h}\right)du\] which can be more easily integrated once the constants are completely separated: \[ F(T) = \frac{2\pi h }{c^2} \left(\frac{kT}{h}\right)^4 \int _0 ^\infty \frac{ u^3 }{e^{u} -1} du,\] and knowing a particular integration of a strange equation, this all becomes much easier \[\int _0 ^\infty \frac{ u^3 }{e^{u} -1} du = \frac{\pi^4}{15}\] \[ F(T) = \frac{2\pi k^4 }{c^2 h^3} T^4 \cdot \frac{\pi^4}{15}.\] As is stated in the problem, all the constants can be grouped together and you would find that we come remarkably close to the actual value of the  Stefan-Boltzmann constant, although I suspect we would get a closer value were we to complete the integration by ourselves, as difficult as that may be. Nevertheless, the constants all become a single term, which simplifies the Bolometric Flux equation into:  \[ F(T)= \sigma T^4\]

b) From the problem, we know we are simply converting the specific intensity according to frequency, to specific intensity according to wavelength. Therefore, if they are both the same value, but in different metrics, they can simply be equaled to the other and then attempt to integrate both, with parameters which leave us with their derivatives with respect to the parameter: \[ -B_\lambda d\lambda = B_\nu d\nu ,\] (we add a negative sign to establish how they are inversely proportional) and by placing all on one side:  \[ -B_\lambda = -B_\nu \frac{d\nu}{d\lambda} ,\]  and after simplifying it down: \[ B_\lambda = B_\nu \frac{d\nu}{d\lambda} ,\] we just find the derivative of \( \nu = \frac{c}{\lambda}\) in terms of \(\lambda\): \[ B_\lambda = B_\nu~ \frac{c}{\lambda^2},\] and so applying it to the original intensity equation (remember to change all \(\nu\)’s into \(c/\lambda\)) : \[ B_\lambda = \frac{2hc^2}{\lambda^5} \frac{1}{e^{hc/\lambda kT} – 1} .\]

c) Now, to maximize the equation, we have to derive the function and set it equal to ):  \[ \frac{d ~B_\lambda}{d\lambda} = \frac{d \frac{2hc^2}{\lambda^5} \frac{1}{e^{hc/\lambda kT} – 1}}{d\lambda} ,\] which can now be interpreted with partial fractions (only deriving with respect to a specific variable: \[\frac{\partial B_\lambda}{\partial d\lambda} = \frac{\partial B\lambda}{\partial u} \frac{\partial u}{\partial \lambda}\] and thus we can define u as \[ u = \frac{hc}{kT\lambda},\] and we place thisn into the equation and get:  \[ B_u = 2hc^2 \frac{\left(\frac{ukT}{hc}\right)^5}{e^u – 1} ,\] and now we take the derivative with respect to u on both  sides and get (with some separation of constants and application of the product rule): \[ \frac{d B_u}{du} = 2hc^2 \left(\frac{kt}{hc}\right)^5  (5u^4(e^u -1 ) – u^5(e^u)),\] so we now take it and optimize it by setting it to 0: \[ 0 = 2hc^2 \left(\frac{kt}{hc}\right)^5  (5u^4(e^u -1 ) – u^5(e^u)),\] we eliminate all the constants and begin solving the equation: \[ 5u^4 (e^u -1 ) = u^5 ~e^u \] \[ 5 = \frac{ue^u}{e^u -1 },\] and now we can make a first order expansion and simplify the equation: \[ 5 = \frac{u e^u}{e^u} \] \[ 5 = u\]

Knowing what we defined to be u: \[ u = \frac{hc}{kT\lambda}\] and solved to find \(\lambda\), we see: \[ \lambda = \frac{hc}{ukT},\] and now we define our constants, the Plank constant: \[ h =6.63 \times 10^{-27} erg \cdot s,\] Boltzmann’s constant: \[ k = 1.38 \times 10^{-16} erg/K,\] and the speed of light:  \[ c = 3 \times 10^{10} cm/s,\] we solve and find that:  \[ \lambda_{max} = \frac{6.63 \times 10^{-27} erg \cdot s \cdot 3 \times 10^{10} cm/s }{ 5 \cdot 1.38 \times 10^{-16} erg/K  \cdot T} \] \[ \lambda_{max} = \frac{0.29 cm \cdot K}{T }\]

d) This part of the problem is simply using a basic equation and expanding it to see how it acts in small frequencies, \[B_\nu(T) = \frac{2\nu^3 h }{c^2 (e^{h\nu / kT} -1)}\] and the expansion is given to us as: \[e^x \approx 1 + x\] which, plugged in, would just be:  \[ B_\nu(T) = \frac{2\nu^3 h }{c^2} \frac{1}{1 + \frac{h\nu}{ kT} -1},\] and thus we are left with: \[ B_\nu(T) = \frac{2\nu^3 h }{c^2} \frac{1}{\frac{h\nu}{ kT}}\]


e) Now, just applying what we learned in part a), we know how\[ F_\nu (T) = \frac{2\pi}{c^2} \frac{\nu^3 h}{e^{h\nu/kT} -1} ,\] can be integrated over all frequencies and we find: \[ F(T) = \sigma T^4,\]then, we have to understand how Flux is distributed isotopically. This means we have to understand the entire area, this being: \[ A = \int^{2\pi}_0 \int^\pi_0 R^2 \sin\theta d\theta d\phi\] \[ A= \int^{2\pi}_0  2R^2 d\phi\] \[ A = 4\pi R^2 \] So by multiplying both sides of the \(F(T)\) equation by the complete area of the light expansion, we have: \[ 4\pi R^2 F = 4\pi R^2 \sigma T^4,\] and Flux in all directions at all times is simply the Luminosity, as such: \[ L = 4\pi R^2 \sigma T^4,\] the relationship between Luminosity and Temperature. 

Monday, February 8, 2016

Gravitational Waves, the Next E-M Spectrum?

We are bounded in a nutshell of Infinite Space: Free Form #2: Gravitational Waves, the Next E-M Spectrum?

In 1915, Albert Einstein published his theory of general relativity, the basis for the future of Astrophysics, although few saw it this way at the time. For decades, general relativity was just a mathematical concept physicists considered enjoyable to wonder at, hardly ever perceiving the importance it would have in describing the universe. General relativity predicted the nature of the Space-Time Fabric, a description of a three-dimensional structure which related time, space, and the interactions of masses and electromagnetism with it. Mass creates gravitational fields, undulations in the fabric of space time which curves it and causes significant effects to light and time. Furthermore, mass con be concentrated in infinitely small spaces, where the density goes to infinity, otherwise known as a black hole, which have severe effects on the universe such as gravitational lensing (like microlensing, if you’ll remember from previous posts, but more drastic), and becoming the center of galaxies, permitting other systems like our planet to eventually develop. 

Moreover, one of the predictions of the General Relativity is the creation of Gravitational Waves (imagine a stone being dropped in a pond, the waves that emerge from that incident are analogous to gravitational waves) from virtually any object with mass. However, these waves are all but nonexistent in what we consider “normal” settings, and as such need to be looked for in other, more elaborate systems which could yield an actual result. Gravitational Waves are propagating tidal forces of gravity which stretch and compact the space-time fabric after it has been exposed to a perturbation, and for it to be noticeable it needs to be something in the range of binary neutron star systems and Black Hole collisions. These events are large enough to create noticeable distortions in the fabric, enough that we may, perhaps, detect them directly. Indirectly however, Hulse and Taylor discovered a binary pulsar star system in 1974 which required the influence of gravitational waves for the rotation model to fit best with the observations, a feat which earned them the 1993 Nobel Prize. 


Nevertheless, in recent years more instruments have been created to accurately measure distortions in Space-time in the order of \(10^{-22} m ,\) at present.  These detectors are called LIGO (Laser Interferometer Gravitational-Wave Observatory) and Advanced LIGO, and by means of a set of mirrors and detectors several kilometers in length, they are designed to detect the smallest possible changes in length which could show evidence of gravitational waves. (Check a 10:40 AM press release from the LIGO team on Thursday, February 11, 2016, to see what they learned.) Moreover, of we were to find gravitational waves, it opens up a completely new field of astronomy observation which can be fine-tuned to effectively measure Black Hole mass, energy outputs, and better understand the fabric of space-time itself. Gravitational waves could very well be as diverse as the E-M spectrum, letting us interpret a completely new way to look at the universe, and slowly come to understand it better. 

References:
Carrol & Ostlie; 2007; pp. 688, 703